Next Article in Journal
DFT Investigations on the Ring-Opening Polymerization of Trimethylene Carbonate Catalysed by Heterocyclic Nitrogen Bases
Next Article in Special Issue
Highly Selective Nitrogen-Doped Graphene Quantum Dots/Eriochrome Cyanine Composite Photocatalyst for NADH Regeneration and Coupling of Benzylamine in Aerobic Condition under Solar Light
Previous Article in Journal
Efficient Cross-Coupling of Acetone with Linear Aliphatic Alcohols over Supported Copper on a Fluorite-Type Pr2Zr2O7
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Doped Carbon Dots Decorated TiO2 Nanorods: A Novel Synthesis Route for Enhanced Photoelectrochemical Water Splitting

1
Department of Chemical Engineering, Yeungnam University, 214-1, Daehak-ro 280, Gyeongsan 712-749, Gyeongbuk-do, Korea
2
Graduate School of Engineering, Department of Polymer Chemistry, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1281; https://doi.org/10.3390/catal12101281
Submission received: 26 September 2022 / Revised: 9 October 2022 / Accepted: 14 October 2022 / Published: 20 October 2022

Abstract

:
Herein, we have successfully prepared self-doped carbon dots with nitrogen elements (NCD) in a simple one-pot hydrothermal carbonization method, using L-histidine as a new precursor. The effect of as-prepared carbon dots was studied for photoelectrochemical (PEC) water splitting by decorating NCDs upon TiO2 nanorods systematically by changing the loading time from 2 h to 8 h (TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h). The successful decorating of NCDs on TiO2 was confirmed by FE-TEM and Raman spectroscopy. The TiO2@NCD4h has shown a photocurrent density of 2.51 mA.cm−2, 3.4 times higher than the pristine TiO2. Moreover, TiO2@NCD4h exhibited 12% higher applied bias photon-to-current efficiency (ABPE) than the pristine TiO2. The detailed IPCE, Mott–Schottky, and impedance (EIS) analyses have revealed the enhanced light harvesting property, free carrier concentration, charge separation, and transportation upon introduction of the NCDs on TiO2. The obtained results clearly portray the key role of NCDs in improving the PEC performance, providing a new insight into the development of highly competent TiO2 and NCDs based photoanodes for PEC water splitting.

1. Introduction

Rapidly spiking global energy demands and pollution caused by the depletion of fossil fuels necessitated the development of natural and renewable sources of energy [1]. Hydrogen is an excellent contender capable of replacing fossil fuels owing to its both eco-friendly and reusable nature. Photoelectrochemical (PEC) water splitting is the most reliable and popular method employed for converting solar light energy into clean and sustainable chemical fuels, such as hydrogen [2,3]. The initial study on photocatalytic water splitting using TiO2 was published way back in 1972 [4]. Since then, different types of semiconductor materials including ZnO, [5] BiVO4, [6] WO3, [7] Fe2O3, [8] SrTiO3, [9] C3N4 [10], and Ta3N [11] were reported as photoelectrodes for PEC. The TiO2 material is considered as the most competent semiconductor for investigating PEC devices due to its characteristics such as advantageous band-edge positions, ease of fabrication, abundance, excellent photocorrosion resistance, eco-friendliness, and cost effective nature [12]. However, application of TiO2 in PEC has been constrained by comparatively greater band gaps for its rutile (3.0 eV) and anatase (3.2 eV) phases [12], severe bulk charge recombination, and slow OER kinetics [13]. As a result, numerous attempts were made to surpass the limitations, such as use of dopants [14], formation of heterojunctions [15], surface modification [16], introduction of defects [17], and quantum dot sensitization [15].
Recently, carbon dots (CDs) have been gaining enormous attention by virtue of their fascinating characteristics such as low cost, simple synthesis, functionalization, superior chemical inertness, and photobleaching resistance. Most essentially, CDs are a viable alternative for heavy metal-based QDs and organic dye, owing to its low toxicity with environmental friendliness [18,19,20]. Since the last decade, astonishing progress has been made in the preparation of CDs either in the top-down or bottom-up route [21,22]. However, new inexpensive, large-scale, and green synthetic approaches of CDs still need to be developed. For instance, a study on the CQDs/BiVO4 and CQDs/NiFe-LDH/BiVO4 demonstrated that after the decoration of CDs on their respective semiconductor, negatively shifted onsite potentials and enhanced charge injection rate were observed in PEC water splitting [22,23,24]. In addition, CDs, such as CQDs/TiO2 [11] CQDs/ZnO [25], CQDs/WO3 [26], CQDs/BiVO4 [1], and CQDs/bFe2O3 [27], etc., can improve the light harvesting nature of photoanode in ultraviolet region and expand the range of visible region.
The CDs decorated TiO2 films have been reported earlier from different origin materials by different methods and utilized as photoanode for PEC [23]. Zhou et al. utilized glucose as precursor and alkali-assisted ultrasonic chemical method to prepare CDs; and spin-coated TiO2 film with CDs solution [15]. Wang et al. employed a hydrothermal method to synthesize CDs from phloroglucinol [24]. Usually, for photo-driven reactions, N-doped carbon dots exhibit improved activity both theoretically and experimentally than CDs, owing to beneficial quantum confinement and were capable of creating defect-rich heterostructures [25,26]. Based on the N-doping source material, light-harvesting ability and energy levels can be modulated [27], while the functionality of NCDs may interpret the interaction with the semi-conducting material [28]. Han et al. described the process of preparation of N-doped CDs (NCDs) anchored to TiO2 photoanode in electrochemical and hydrothermal methods by using graphite rods and ammonia to obtain a nitrogen-doped CDs (NCDs) solution. This report has demonstrated the enhanced PEC efficiency due to an increased interface charge transfer [12,29]. The report by Wang et al. on NCDs@TiO2 showed an enhanced photocatalytic property owing to its extended light responses with narrowed bandgap upon introduction of NCDs [30]. However, due to the complexity of NCDs with regards to energy states and chemical structure, the mechanism of NCDs in boosting PEC performance remains unknown [25,31]. Moreover, synthesis of CDs and preparation of photoanode was proceeded in multiple steps, which again increases the preparation cost of the electrode [32,33]. Therefore, it is of critical importance to prepare at low cost, as well as understand the nanostructure of NCDs, their interfacial interactions with semiconductor materials and further developments of NCDs.
In the present study, we report the synthesis of new NCDs decorated TiO2 film in a simple one-pot hydrothermal method using L-histidine as source material. The effect of NCDs on TiO2 nanorod film for PEC water splitting has been analyzed systematically by changing the NCDs’ loading time from 2 h–8 h. The prepared photoanodes are named as TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h. NCDs loaded photoanodes showed higher PEC performance than pristine TiO2, suggesting the contribution of NCDs towards enhancing the performance of PEC. The highest efficiency was found for TiO2@NCD4h (2.51 mA.cm−2), 3.4 times greater than pristine TiO2 (0.73 mA.cm−2). The higher photocurrent for TiO2@NCDs could be ascribed to the improved light harvesting property, decreased rate of recombination, and increased charge carrier density. The detailed characterization of NCDs and NCD loaded TiO2 and PEC water splitting performance analysis were performed and discussed.

2. Results and Discussion

2.1. Characterization

FE-SEM and HR-TEM analyses were executed in order to assess the successful loading of NCDs on TiO2 and their morphology and the obtained images are illustrated in Figure 1 and Figure 2. The FE-SEM analysis of pristine TiO2 film (Figure 1a) has shown dense nanorod morphology of TiO2 which have perpendicularly grown on FTO glass showing an average length of ~2.8 μm and width of ~150 nm. Moreover, no obvious changes in the size and morphology of TiO2 were observed in Figure 1b, even after dipping for 8 h in NCDs solution. Further, HR-TEM (Figure 1c) analysis confirmed the nanorod morphology of TiO2. Moreover, the observed lattice fringes’ distance in Figure 1d was 0.35 nm, which corresponds to the d-spacings of the rutile TiO2 (101) planes, which has well-matched with XRD results [12]. Further, HRTEM image of TiO2@NCD4h (Figure 2) showed that NCDs are uniformly loaded on the TiO2 nanorods and appeared in a sphere and ellipsoidal morphology with particle size ranging from 4 to 10 nm. In addition, 0.21 nm lattice spacing was observed for NCDs particle, associated with the (100) facet of NCDs (Figure S3) [15,18,34]. Moreover, to further investigate the distribution of Ti, O, C, and N elements, the elemental mapping analysis was executed, and the respective results are displayed in Figure 2b. The obtained results have shown even distribution of C and N elements on TiO2 nanorods’ surface, suggesting the successful decoration of NCDs on the TiO2.
The crystalline structure of the as-synthesized TiO2 and the effect of NCDs loading time on TiO2 crystallinity (TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h) and the orientation growth were examined using XRD analysis. The obtained XRD peaks are displayed in Figure 3. The diffraction peaks of pristine TiO2 films appeared at 36.10°, 41.27°, 54.39°, 62.86°, and 69.80° and correspond to the (101), (111), (211), (002), and (112) crystal planes of tetragonal rutile structure [15,18,35,36]. Moreover, regardless of the loading time of NCDs, the peak positions are the same, but the (101) plane intensity has increased with NCDs loading time. The results suggest that the TiO2 nanorod crystal structure does not get affected by loading NCDs but size of the crystal and preferred orientation directions sparsely get affected. Furthermore, no noticeable diffraction peak of NCDs was observed for TiO2-NCDs, which could be attributed to the modest load of NCDs, lower than the minimum limitation of XRD detection [15,35].
As seen in Figure 4, the Raman peaks of TiO2 located at 615.2, 450.5, and 240 cm−1 correspond to (A1g), (Eg), and multi-photon scattering process, respectively, and represent the TiO2 rutile phase. The peaks which appeared at 1580 and 1333 cm−1 can be attributed to the D (disordered sp2) band and G band of NCDs, respectively [37,38]. Thus, the Raman spectrum of TiO2@NCD4h has shown five peaks, indicating successful fabrication of NCDs in TiO2 nanorods. Furthermore, the enhancement of Raman intensity might be contributed by the increased crystallinity, and it is consistent with the XRD results.
The elemental composition and chemical binding of NCDs decorated on TiO2 catalyst were determined by the XPS analysis (Figure 5). The survey scans illustrated the elements in two structures, e.g., C, N, Ti, and O elements in TiO2@NCDs4h, whereas C, Ti, and O elements were in the pristine TiO2 [12,39] (Figure 5a). An increase in the carbon content and the presence of N1s peak compared with the bare TiO2 evidently confirm that the NCDs were successfully decorated on TiO2. Ti2p spectra (Figure 5b) showed two representative peaks at 464.0 and 458.4 eV (difference: 5.6 eV), which correspond to the spin-orbit coupling for Ti2p1/2and Ti2p3/2, respectively, and was identical to those for TiO2 [15]. Furthermore, Ti2p peaks of the TiO2@NCDs4h structure have considerably shifted (by 0.2 eV) compared to bare TiO2 (Figure 5b). This is due to the electronegativity of C/Ns, which increased the binding ability of extra-nuclear electrons, hence raising the binding energy. The C 1s spectra (Figure 5c) is fitted with three peaks corresponding to C-C, C-O & C-N, and C=O & C=N bonds at 284.8, 286.0, and 288.3 eV, respectively [40]. The N1s spectra (Figure 5d) shows three peaks which appeared at 396.7 eV, 401.4 eV, and 403.7 eV, ascribed to the pyridine-N, pyrrole-N, and Graphitic-N, respectively, indicating the carbon dot doped with the N element [41,42]. In Figure S4, the peak in the O1 s region of TiO2@NCDs@4h was deconvoluted into three peaks at 532.3, 530.1, and 529.6 eV, which are assigned to O-H, C-O, or O-N, and Ti-O bonds in TiO2@NCDs4h, respectively. In agreement with the above microstructure analysis, XPS results suggest that NCDs were successfully deposited on the TiO2 [40,42,43].
The absorption properties of the materials are important parameters to estimate the light harvesting nature and energy levels to be used in solar energy conversions. The optical properties of the prepared NCDs in solution and on TiO2 photoanode with changing time were analyzed systematically. The UV-vis spectra of NCDs in solution and TiO2@NCDs thin films are shown in Figure 6a,b and c by changing the loading time. The absorption band of pristine TiO2 ~400 nm represents the rutile TiO2 band edge [44]. After introduction of the NCDs on TiO2, the light harvesting property was enhanced with the increased amount of NCDs loading. The results suggest the successful decoration of NCDs and their contribution in improving the light harvesting property. Moreover, the Tauc plot Equation (S1) was employed to calculate the bandgap (Eg) of pristine TiO2 and TiO2@NCD photoanodes [36,45,46]. The calculated Eg values of TiO2, TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h were 3.12, 3.07, 3.03, 3.04, and 3.05 eV, respectively.

2.2. PEC Performance of the Photoanodes

The effect of newly prepared NCDs on photoelectrochemical water oxidation was studied systematically by changing the loading time of NCDs on TiO2 film and comparing it with the pristine TiO2. Figure 7a shows linear sweep voltammetry performance of TiO2, TiO2@2hNCD, TiO2@4hNCDs, TiO2@6h NCDs, and TiO2@8hNCDs, and their photocurrent data at 1.23 V are illustrated in Table S2. The pristine TiO2 film displayed a 0.73 mA.cm−2 photocurrent at 1.89 V vs. RHE. After loading NCDs upon the TiO2 film, an enhanced photocurrent was observed compared the pristine TiO2, which instigated to perform optimization studies to improve the NCDs loading and thereby achieve optimum PEC performance using TiO2 with NCDs. In order to optimize the TiO2@NCDs photoanodes, the decorated NCDs on TiO2 was controlled by monitoring the hydrothermal reaction. In Figure 7a, the photocurrents of four NCDs decorated TiO2 (TiO2@NCDs)-based photoanodes showed higher photocurrent than pristine TiO2, indicating the contribution of NCDs in enhancing PEC performance of TiO2. Significantly, the photoanode corresponding to TiO2@NCD2h has displayed an improved photocurrent density of 2.33 mA.cm−2 at 1.89 V vs. RHE, while pristine TiO2 photoanode has shown 0.73 mA.cm−2 at 1.89 V vs. RHE. Further, by increasing the loading time from 2 h to 4 h (TiO2@NCD4h), the photocurrent density has also increased to 2.51 mA.cm−2 at 1.89 V vs. RHE, which was 3.4 times greater than the pristine TiO2. Moreover, TiO2@NCD4h photoanode possesses both enhanced photocurrent density and smaller onset potential than pristine TiO2. The higher photoresponse of NCD decorated TiO2 might be due to the addition of NCD which could effectively promote the separation of photogenerated electron-hole, and promote the capture of water molecules and intermediates in the process of water decomposition by electrons and holes at the interfaces [15,47,48]. However, further increasing the NCDs loading by increasing loading time to 6 h (TiO2@NCD6h) and 8 h (TiO2@NCD8h) showed declined photocurrent density of 1.99 and 1.85 mA.cm−2 at 1.89 V vs. RHE, respectively. In addition, onsite potentials also increased compared to the TiO2@NCD4h-based films, possibly due to the variation in conductivity by decorated NCDs. This phenomenon will be further discussed in electrochemical impedance spectroscopy (EIS) section [29].
Chronoamperometric analysis was performed under chopped illumination for all the prepared electrodes at 1.89 V vs. RHE for 30 s to better understand photo response with time and stability. As depicted in Figure 7b, the photocurrent has rapidly increased immediately after illumination and sharply fell to zero upon stopping the illumination. It confirms that the prepared photoanodes have a well-reproducible photocurrent. Meanwhile, after decorating NCDs on TiO2, the response speed was boosted compared with the pristine TiO2, indicating that the presence of NCDs can significantly reduce the charge recombination in the intersection of electrolyte and photoanode surface. The highest photocurrent density was detected for the TiO2@NCD4h, which is consistent with the observed LSV results. In order to understand the durability of the prepared electrode, 1 h of continuous illuminated chronoamperometric analysis was performed with a high performing photoanode (TiO2@NCD4h) by comparing with the pristine TiO2 based photo anode at same experimental condition. As displayed in Figure 7c, after continuous illumination for 1 h, TiO2@NCD4h has retained 99% of its initial activity, which supports the excellent stability of NCDs decorated TiO2 photo anode. The current densities have matched well with the LSV data. Both TiO2 and TiO2@NCD4h showed excellent stability after 1 h of continuous irradiation without any photocurrent decay.
The incident photon-to-current conversion efficiency (IPCE) was evaluated to analyze the contribution of various photons in obtaining solar photocurrent. The IPCE has been deduced by using the formula (1):
IPCE   ( % ) = 1240 J ( λ ) λ P ( λ ) × 100   ( % )
where, P(λ), λ, and J(λ) are the intensity of a specific wavelength, wavelength of incident light, and photocurrent density at specific wavelength, respectively. Figure 8a shows enhanced IPCE after decorating NCDs on TiO2 with highest IPCE of 29.76% at ~390 nm for TiO2@NCD4h, which was ~3 times greater than the pristine TiO2 (9.76%). The IPCE trend was consistent with the obtained photocurrent density and the enhanced IPCE region is in good agreement with the optical absorption properties. The improved IPCE after the introduction of NCDs to the TiO2 reveals the contribution of NCDs in obtaining enhanced photocurrent density.
Besides, the applied bias photon-to-current efficiency (ABPE) has been calculated by using the Equation (2):
ABPE   ( % ) = J ( 1.23 V ) P × 100   ( % )
where P, J, and V are the power density of incident light (100 mW cm−2), photocurrent density (mA cm−2), and the applied bias (V vs. RHE), respectively. As seen in Figure 8b, the pristine TiO2 reached maximum 0.11% ABPE at 0.88 V vs. RHE, while TiO2@NCD4h reached 1.37% photo conversion efficiency at 0.36 V vs. RHE, 12 times greater than the pristine TiO2 ABPE, suggesting an effective electron-hole pairs separation after the introduction of NCDs [12].
To further comprehend the interfacial charge transfer kinetics at the intersection of photoanode and electrolyte, EIS was employed under illumination and the respective Nyquist plots are displayed in Figure 9a. The decreased radius order of semi-circle was TiO2 > TiO2@NCD8h > TiO2@NCD6h > TiO2@NCD2h > TiO2@NCD4h. The smallest arc of the TiO2@NCD4h compared with its counter parts demonstrates the improved interfacial charge transfer kinetics due to the introduction of NCDs [18]. Furthermore, using Zview software program, the EIS curves have been fitted with an analogous circuit model given in Figure 9a inset, where CPE, Rct, and Rs indicate the constant phase element, charge transfer resistance, and series resistance at the electrolyte/electrode interface, respectively. The observed Rct values of TiO2 and NCDs decorated TiO2 films (2 h–8 h) were 376.0 Ω, 273.6 Ω, 248.3 Ω, 277.2 Ω, and 297.4 Ω, respectively. The lowest Rct value of TiO2@NCD4h further demonstrates the advantage of NCDs decorated TiO2 nanorods in enhancing the charge separation and transfer kinetics.
Mott–Schottky analyses have been executed to estimate the energy band position of pristine TiO2 and TiO2@NCDs, and the corresponding curves are displayed in Figure 9b and the data are depicted in Table S1. The positive slope of both curves indicates n-type semiconductor of TiO2 [18]. The flat band (VFB) potential can be calculated following the Equation (S1). The obtained VFB values of TiO2 and TiO2@NCDs (2 h, 4 h, 6 h, and 8 h) were 0.17, 0.27, 0.34, 0.22, and 0.19 V vs. NHE, respectively, which could be accomplished by the X-axis intercept. Moreover, NCDs decorated TiO2 films have shown decreased VFB than pristine TiO2, suggesting an increased band bending of the photoanode, favorable to enhance the charge transfer between the photoanode interfaces and electrolyte. Eventually, enhanced PEC was observed for NCDs decorated TiO2 films. As per the available literature [49,50], the bottom of the conduction band (CB) was −0.1 V lower than the VFB of an n-type semiconductor [50]. Therefore, the CB of TiO2 and TiO2@NCDs were estimated to be positioned at less than 0.1 V of their VFB [33]. Based on the VFB, the CB edge of TiO2 and TiO2@NCDs (2 h, 4 h, 6 h, and 8 h) were determined to be at 0.07, 0.17, 0.24, 0.12, and 0.09 V, respectively. Particularly, doping of NCDs promotes a downward shift in energy levels towards higher potentials and enhances the carrier density in TiO2 [51].

3. Experimental

3.1. Materials

All chemicals were used directly without purifying any further. Hydrochloric acid (35%) was obtained from OCI Company Ltd., titanium butoxide (TBOT, 98%) was purchased from Sigma-Aldrich, L-histidine (98%) was procured from Alfa Aesar, and sodium sulfate anhydrous (99%) was obtained from Duksan. For all the experiments, Milli-Q water (MΩ 18) was used.

3.2. Preparation of Rutile TiO2 Film (TiO2):

The FTO coated glasses (1.5 mm × 2.5 mm, 8 Ω/cm2) were cleaned ultrasonically using detergent, milli-Q water, ethanol, and acetone for 1 h, respectively. TiO2 film was synthesized by following a reported hydrothermal method with certain modifications [18]. Under continuous stirring, 0.33 mL titanium (IV) butoxide was added dropwise to 20 mL equal volumes of HCl (35%) and milli-Q water mixed solution until it turned translucent. The solution was then moved to a 50 mL autoclave lined with Teflon, and the FTO glass was placed against the walls of Teflon vessel, conducted side down, for 12 h and heated to 150 °C. The TiO2 layer was completely cleaned with milli-Q water and ethanol after cooling to RT, before being sintered in air at 450 °C for 1 h.

3.3. Preparation of TiO2@NCDs:

NCDs have been prepared using a hydrothermal approach (Scheme 1). First, 0.2 g of L-histidine was included in 20 mL mixture of milli-Q water and HCl in the ratio of 19:1. The solution was shifted to 50 mL autoclave, and two TiO2 films on FTO glasses were inserted in the Teflon vessel with the TiO2 side facing down. Then, the hydrothermal reaction was performed at 180 °C for 2, 4, 6, 8, and 10 h. The samples were denoted as TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, TiO2@NCD8h, respectively (Figures S1 and S2). The TiO2@NCD films were extensively washed with milli-Q H2O and ethanol upon cooling to RT. Then, copper wires and as prepared photoanodes were adhered using silver paint. The samples were air dried for 3 h. Finally, the samples were encased by nonconductive epoxy with the illuminated area of 1 cm2 and left to rest in air for at least 3 h.

4. Conclusions

In conclusion, the new NCDs were successfully prepared in a simple one-pot hydrothermal synthesis method using L-histidine as an initial precursor and the as-prepared NCDs were decorated on the TiO2 nanorod-based photoanode. The as-prepared NCDs and NCD decorated TiO2 nanorods were well characterized using FE-SEM, HR-TEM, EDS elemental mapping, which revealed the nanorod morphology of TiO2 and uniform distribution of CDs on TiO2 surface while, XPS and Raman analyses have confirmed the successful self nitrogen element doping, preparation and decorating of NCDs on TiO2 nanorod. The effect of NCDs decorated TiO2 was tested for photoelectrochemical water splitting analysis systematically by changing the loading time of NCDs from 2 h to 8 h. The highest efficiency was observed for the TiO2@NCD4h-based photoanode (2.51 mA.cm−2), which was a 3.4 times higher photocurrent density than the pristine TiO2-based photoanode. It might be attributed to the increased light harvesting property with charge separation and transportation. The observed IPCE of TiO2@NCD4h has shown 3 times higher quantum yield (29.76%) than pristine TiO2 (9.76). In addition, the calculated ABPE was 12% higher for TiO2@NCD4h than the pristine TiO2, which revealed the enhanced light harvesting property of photoanodes upon loading the NCDs. Moreover, the reduced charge transfer resistance and higher charge carrier density, as observed from EIS and Mott–Schottky analyses, respectively, further support the advantage of newly prepared NCDs in enhancing the PEC performance by promoting effective charge separation and transportation. This study may open up new insights into the rational design and synthesis of highly efficient photoanodes for PEC water splitting.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12101281/s1; Figure S1: Imagies of NCDs’ solution under natural and UV light. Figure S2. NCDs solution based on the reaction time. Figure S3. TEM of NCD. Figure S4. XPS O 1s spectra of TiO2@NCDs4h. Table S1. EIS Data of TiO2 and NCD decorated TiO2 photoanodes Table S2. The photocurrent densities of the TiO2 and TiO2@NCDs photoanodes.

Author Contributions

Conceptualization, G.K., C.T.T.T.; Experiments, C.T.T.T., G.S., L.J.; data curation, C.T.T.T., G.K.; formal analysis, H.D.K.; funding acquisition, J.H.K.; supervision, J.H.K. and G.K. writing—review and editing, C.T.T.T., G.K. and J.H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (20214000000720). This work was supported by “Human Resources Program in Energy Technology” of the Korea Institute of Energy Technology Evaluation and Planning (KETEP), granted financial resource from the Ministry of Trade, Industry & Energy, Republic of Korea. (No. 20204010600100).

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, Q.; Guo, B.; Yu, J.; Ran, J.; Zhang, B.; Yan, H.; Gong, J.R. Highly efficient visible-light-driven photocatalytic hydrogen production of CdS-cluster-decorated graphene nanosheets. J. Am. Chem. Soc. 2011, 133, 10878–10884. [Google Scholar] [CrossRef] [PubMed]
  2. Li, H.; Yu, H.; Sun, L.; Zhai, J.; Han, X. A self-assembled 3D Pt/TiO2 architecture for high-performance photocatalytic hydrogen production. Nanoscale 2015, 7, 1610–1615. [Google Scholar] [CrossRef] [PubMed]
  3. Alshorifi, F.T.; Ali, S.L.; Salama, R.S. Promotional Synergistic Effect of Cs–Au NPs on the Performance of Cs–Au/MgFe2O4 Catalysts in Catalysis 3, 4-Dihydropyrimidin-2 (1H)-Ones and Degradation of RhB Dye. J. Inorg. Organomet. Polym. Mater. 2022, 32, 3765–3776. [Google Scholar] [CrossRef]
  4. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, W.-D.; Jiang, L.-C.; Ye, J.-S. Photoelectrochemical study on charge transfer properties of ZnO nanowires promoted by carbon nanotubes. J. Phys. Chem. C 2009, 113, 16247–16253. [Google Scholar] [CrossRef]
  6. Kim, T.W.; Ping, Y.; Galli, G.A.; Choi, K.-S. Simultaneous enhancements in photon absorption and charge transport of bismuth vanadate photoanodes for solar water splitting. Nat. Commun. 2015, 6, 8769. [Google Scholar] [CrossRef] [Green Version]
  7. Hou, Y.; Zuo, F.; Dagg, A.P.; Liu, J.; Feng, P. Branched WO3 nanosheet array with layered C3N4 heterojunctions and CoOx nanoparticles as a flexible photoanode for efficient photoelectrochemical water oxidation. Adv. Mater. 2014, 26, 5043–5049. [Google Scholar] [CrossRef]
  8. Liu, S.; Zheng, L.; Yu, P.; Han, S.; Fang, X. Novel composites of α-Fe2O3 tetrakaidecahedron and graphene oxide as an effective photoelectrode with enhanced photocurrent performances. Adv. Funct. Mater. 2016, 26, 3331–3339. [Google Scholar] [CrossRef]
  9. Liang, Z.; Hou, H.; Song, K.; Zhang, K.; Fang, Z.; Gao, F.; Wang, L.; Chen, D.; Yang, W.; Zeng, H. Boosting the photoelectrochemical activities of all-inorganic perovskite SrTiO3 nanofibers by engineering homo/hetero junctions. J. Mater. Chem. A 2018, 6, 17530–17539. [Google Scholar] [CrossRef]
  10. Ye, L.J.; Wang, D.; Chen, S.J. Fabrication and Enhanced Photoelectrochemical Performance of MoS2/S-Doped g-C3N4 Heterojunction Film. ACS Appl. Mater. Interfaces 2016, 8, 5280–5289. [Google Scholar] [CrossRef]
  11. Liu, G.; Fu, P.; Zhou, L.; Yan, P.; Ding, C.; Shi, J.; Li, C. Efficient Hole Extraction from a Hole-Storage-Layer-Stabilized Tantalum Nitride Photoanode for Solar Water Splitting. Chemistry 2015, 21, 9624–9628. [Google Scholar] [CrossRef] [PubMed]
  12. Liang, Z.; Hou, H.; Fang, Z.; Gao, F.; Wang, L.; Chen, D.; Yang, W. Hydrogenated TiO2 Nanorod Arrays Decorated with Carbon Quantum Dots toward Efficient Photoelectrochemical Water Splitting. ACS Appl. Mater. Interfaces 2019, 11, 19167–19175. [Google Scholar] [CrossRef] [PubMed]
  13. Zhou, T.; Chen, S.; Wang, J.; Zhang, Y.; Li, J.; Bai, J.; Zhou, B. Dramatically enhanced solar-driven water splitting of BiVO4 photoanode via strengthening hole transfer and light harvesting by co-modification of CQDs and ultrathin β-FeOOH layers. Chem. Eng. J. 2021, 403, 126350. [Google Scholar] [CrossRef]
  14. Wang, D.-H.; Jia, L.; Wu, X.-L.; Lu, L.-Q.; Xu, A.-W. One-step hydrothermal synthesis of N-doped TiO2/C nanocomposites with high visible light photocatalytic activity. Nanoscale 2012, 4, 576–584. [Google Scholar] [CrossRef]
  15. Zhou, T.S.; Chen, S.; Li, L.S.; Wang, J.C.; Zhang, Y.; Li, J.H.; Bai, J.; Xia, L.G.; Xu, Q.J.; Rahim, M.; et al. Carbon quantum dots modified anatase/rutile TiO2 photoanode with dramatically enhanced photoelectrochemical performance. Appl. Catal. B-Environ. 2020, 269, 118776. [Google Scholar] [CrossRef]
  16. Wen, P.; Su, F.J.; Li, H.; Sun, Y.H.; Liang, Z.Q.; Liang, W.K.; Zhang, J.C.; Qin, W.; Geyer, S.M.; Qiu, Y.J.; et al. A Ni2P nanocrystal cocatalyst enhanced TiO2 photoanode towards highly efficient photoelectrochemical water splitting. Chem. Eng. J. 2020, 385, 123878. [Google Scholar] [CrossRef]
  17. Cheng, X.; Dong, G.; Zhang, Y.; Feng, C.; Bi, Y. Dual-bonding interactions between MnO2 cocatalyst and TiO2 photoanodes for efficient solar water splitting. Appl. Catal. B: Environ. 2020, 267, 118723. [Google Scholar] [CrossRef]
  18. Zhou, T.; Li, L.; Li, J.; Wang, J.; Bai, J.; Xia, L.; Xu, Q.; Zhou, B. Electrochemically reduced TiO2 photoanode coupled with 426 oxygen vacancy-rich carbon quantum dots for synergistically improving photoelectrochemical performance. Chem. Eng. J. 2021, 425, 131770. [Google Scholar] [CrossRef]
  19. Wang, H.-J.; Yu, T.-T.; Chen, H.-L.; Nan, W.-B.; Xie, L.-Q.; Zhang, Q.-Q. A self-quenching-resistant carbon dots powder with tunable solid-state fluorescence and their applications in light-emitting diodes and fingerprints detection. Dye. Pigment. 2018, 159, 245–251. [Google Scholar] [CrossRef]
  20. Tangy, A.; Kumar, V.B.; Pulidindi, I.N.; Kinel-Tahan, Y.; Yehoshua, Y.; Gedanken, A. In-situ transesterification of Chlorella vulgaris using carbon-dot functionalized strontium oxide as a heterogeneous catalyst under microwave irradiation. Energy Fuels 2016, 30, 10602–10610. [Google Scholar] [CrossRef]
  21. Feng, Z.; Adolfsson, K.H.; Xu, Y.; Fang, H.; Hakkarainen, M.; Wu, M. Carbon dot/polymer nanocomposites: From green synthesis to energy, environmental and biomedical applications. Sustain. Mater. Technol. 2021, 29, e00304. [Google Scholar] [CrossRef]
  22. Beutier, C.; Serghei, A.; Cassagnau, P.; Heuillet, P.; Cantaloube, B.; Selles, N.; Morfin, I.; Sudre, G.; David, L.J.P. In situ coupled mechanical/electrical/WAXS/SAXS investigations on ethylene propylene diene monomer resin/carbon black nanocomposites. Polymer 2022, 254, 125077. [Google Scholar] [CrossRef]
  23. Sendão, R.M.S.; Esteves da Silva, J.C.G.; Pinto da Silva, L. Photocatalytic removal of pharmaceutical water pollutants by TiO2—Carbon dots nanocomposites: A review. Chemosphere 2022, 301, 134731. [Google Scholar] [CrossRef]
  24. Wang, X.; Wang, M.; Liu, G.; Zhang, Y.; Han, G.; Vomiero, A.; Zhao, H. Colloidal carbon quantum dots as light absorber for 444 efficient and stable ecofriendly photoelectrochemical hydrogen generation. Nano Energy 2021, 86, 106122. [Google Scholar] [CrossRef]
  25. Luo, H.; Dimitrov, S.; Daboczi, M.; Kim, J.-S.; Guo, Q.; Fang, Y.; Stoeckel, M.-A.; Samorì, P.; Fenwick, O.; Jorge Sobrido, A.B.; et al. Nitrogen-Doped Carbon Dots/TiO2 Nanoparticle Composites for Photoelectrochemical Water Oxidation. ACS Appl. Nano Mater. 2020, 3, 3371–3381. [Google Scholar] [CrossRef]
  26. Hola, K.; Sudolská, M.; Kalytchuk, S.; Nachtigallová, D.; Rogach, A.L.; Otyepka, M.; Zboril, R. Graphitic nitrogen triggers red fluorescence in carbon dots. ACS Nano 2017, 11, 12402–12410. [Google Scholar] [CrossRef] [PubMed]
  27. Hu, R.; Li, L.; Jin, W.J. Controlling speciation of nitrogen in nitrogen-doped carbon dots by ferric ion catalysis for enhancing fluorescence. Carbon 2017, 111, 133–141. [Google Scholar] [CrossRef]
  28. Xie, S.; Su, H.; Wei, W.; Li, M.; Tong, Y.; Mao, Z. Remarkable photoelectrochemical performance of carbon dots sensitized TiO2 under visible light irradiation. J. Mater. Chem. A 2014, 2, 16365–16368. [Google Scholar] [CrossRef]
  29. Han, Y.; Wu, J.; Li, Y.; Gu, X.; He, T.; Zhao, Y.; Huang, H.; Liu, Y.; Kang, Z. Carbon dots enhance the interface electron transfer and photoelectrochemical kinetics in TiO2 photoanode. Appl. Catal. B: Environ. 2022, 304, 120983. [Google Scholar] [CrossRef]
  30. Wang, Q.; Cai, J.; Biesold-McGee, G.V.; Huang, J.; Ng, Y.H.; Sun, H.; Wang, J.; Lai, Y.; Lin, Z. Silk fibroin-derived nitrogen-doped carbon quantum dots anchored on TiO2 nanotube arrays for heterogeneous photocatalytic degradation and water splitting. Nano Energy 2020, 78, 105313. [Google Scholar] [CrossRef]
  31. Tian, J.; Leng, Y.; Zhao, Z.; Xia, Y.; Sang, Y.; Hao, P.; Zhan, J.; Li, M.; Liu, H. Carbon quantum dots/hydrogenated TiO2 nanobelt heterostructures and their broad spectrum photocatalytic properties under UV, visible, and near-infrared irradiation. Nano Energy 2015, 11, 419–427. [Google Scholar] [CrossRef]
  32. Ning, X.; Huang, J.; Li, L.; Gu, Y.; Jia, S.; Qiu, R.; Li, S.; Kim, B.H. Homostructured rutile TiO2 nanotree arrays thin film electrodes with nitrogen doping for enhanced photoelectrochemical performance. J. Mater. Sci. Mater. Electron. 2019, 30, 16030–16040. [Google Scholar] [CrossRef]
  33. Altass, H.M.; Morad, M.; Khder, A.E.-R.S.; Mannaa, M.A.; Jassas, R.S.; Alsimaree, A.A.; Ahmed, S.A.; Salama, R.S. Enhanced catalytic activity for CO oxidation by highly active Pd nanoparticles supported on reduced graphene oxide/copper metal organic framework. J. Taiwan Inst. Chem. Eng. 2021, 128, 194–208. [Google Scholar] [CrossRef]
  34. Liu, Y.; Wang, J.; Wu, J.; Zhao, Y.; Huang, H.; Liu, Y.; Kang, Z. Critical roles of H2O and O2 in H2O2 photoproduction over biomass derived metal-free catalyst. Appl. Catal. B: Environ. 2022, 319, 121944. [Google Scholar] [CrossRef]
  35. Masuda, Y.; Kato, K. Synthesis and phase transformation of TiO2 nano-crystals in aqueous solutions. J. Ceram. Soc. Jpn. 2009, 117, 373–376. [Google Scholar] [CrossRef] [Green Version]
  36. Hu, A.; Zhang, X.; Luong, D.; Oakes, K.D.; Servos, M.R.; Liang, R.; Kurdi, S.; Peng, P.; Zhou, Y. Adsorption and Photocatalytic Degradation Kinetics of Pharmaceuticals by TiO2 Nanowires During Water Treatment. Waste Biomass Valorization 2012, 3, 443–449. [Google Scholar] [CrossRef]
  37. Hanaor, D.A.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  38. Ma, H.L.; Yang, J.Y.; Dai, Y.; Zhang, Y.B.; Lu, B.; Ma, G.H. Raman study of phase transformation of TiO2 rutile single crystal irradiated by infrared femtosecond laser. Appl. Surf. Sci. 2007, 253, 7497–7500. [Google Scholar] [CrossRef]
  39. Jiang, D.; Xu, Y.; Hou, B.; Wu, D.; Sun, Y. Synthesis of visible light-activated TiO2 photocatalyst via surface organic modification. J. Solid State Chem. 2007, 180, 1787–1791. [Google Scholar] [CrossRef]
  40. Wei, N.; Liu, Y.; Feng, M.; Li, Z.; Chen, S.; Zheng, Y.; Wang, D. Controllable TiO2 core-shell phase heterojunction for efficient photoelectrochemical water splitting under solar light. Appl. Catal. B: Environ. 2019, 244, 519–528. [Google Scholar] [CrossRef]
  41. Yang, J.; Bai, H.; Tan, X.; Lian, J. IR and XPS investigation of visible-light photocatalysis—Nitrogen–carbon-doped TiO2 film. Appl. Surf. Sci. 2006, 253, 1988–1994. [Google Scholar] [CrossRef]
  42. Wu, D.; Zhang, W.; Feng, Y.; Ma, J. Necklace-like carbon nanofibers encapsulating V3S4 microspheres for ultrafast and stable potassium-ion storage. J. Mater. Chem. A 2020, 8, 2618–2626. [Google Scholar] [CrossRef]
  43. Song, J.; Zheng, M.; Yuan, X.; Li, Q.; Wang, F.; Ma, L.; You, Y.; Liu, S.; Liu, P.; Jiang, D.; et al. Electrochemically induced Ti3+ self-doping of TiO2 nanotube arrays for improved photoelectrochemical water splitting. J. Mater. Sci. 2017, 52, 6976–6986. [Google Scholar] [CrossRef]
  44. Zhuang, H.; Zhang, S.; Lin, M.; Lin, L.; Cai, Z.; Xu, W. Controlling interface properties for enhanced photocatalytic performance: A case-study of CuO/TiO2 nanobelts. Mater. Adv. 2020, 1, 767–773. [Google Scholar] [CrossRef]
  45. Koyyada, G.; Goud, B.S.; Devarayapalli, K.C.; Shim, J.; Vattikuti, S.P.; Kim, J.H. BiFeO3/Fe2O3 electrode for photoelectrochemical water oxidation and photocatalytic dye degradation: A single step synthetic approach. Chemosphere 2022, 303, 135071. [Google Scholar] [CrossRef]
  46. Mohamad, M.; Ul Haq, B.; Ahmed, R.; Shaari, A.; Ali, N.; Hussain, R. A density functional study of structural, electronic and optical properties of titanium dioxide: Characterization of rutile, anatase and brookite polymorphs. Mater. Sci. Semicond. Process. 2015, 31, 405–414. [Google Scholar] [CrossRef]
  47. Wang, T.; Long, X.; Wei, S.; Wang, P.; Wang, C.; Jin, J.; Hu, G. Boosting Hole Transfer in the Fluorine-Doped Hematite Photoanode by Depositing Ultrathin Amorphous FeOOH/CoOOH Cocatalysts. ACS Appl. Mater. Interfaces 2020, 12, 49705–49712. [Google Scholar] [CrossRef]
  48. Fan, X.; Gao, B.; Wang, T.; Huang, X.; Gong, H.; Xue, H.; Guo, H.; Song, L.; Xia, W.; He, J. Layered double hydroxide modified WO3 nanorod arrays for enhanced photoelectrochemical water splitting. Appl. Catal. A: Gen. 2016, 528, 52–58. [Google Scholar] [CrossRef]
  49. Goud, B.S.; Koyyada, G.; Jung, J.H.; Reddy, G.R.; Shim, J.; Nam, N.D.; Vattikuti, S.P. Surface oxygen vacancy facilitated Z-scheme MoS2/Bi2O3 heterojunction for enhanced visible-light driven photocatalysis-pollutant degradation and hydrogen production. Int. J. Hydrogen Energy 2020, 45, 18961–18975. [Google Scholar] [CrossRef]
  50. Chen, W.-Q.; Li, L.-Y.; Li, L.; Qiu, W.-H.; Tang, L.; Xu, L.; Xu, K.-J.; Wu, M.-H. MoS2/ZIF-8 hybrid materials for environmental catalysis: Solar-driven antibiotic-degradation engineering. Engineering 2019, 5, 755–767. [Google Scholar] [CrossRef]
  51. Kalanur, S.S. Structural, Optical, Band Edge and Enhanced Photoelectrochemical Water Splitting Properties of Tin-Doped WO3. Catalysts 2019, 9, 456. [Google Scholar] [CrossRef]
Figure 1. (a,b) Typical SEM of TiO2, TiO2@NCD4h. The corresponding cross-sectional SEM images are shown in the insets. (c) FETEM images of TiO2. (d) FETEM images of TiO2@NCD4h.
Figure 1. (a,b) Typical SEM of TiO2, TiO2@NCD4h. The corresponding cross-sectional SEM images are shown in the insets. (c) FETEM images of TiO2. (d) FETEM images of TiO2@NCD4h.
Catalysts 12 01281 g001
Figure 2. (a1a3) FETEM of TiO2@NCD4h. (b) HAADF-STEM of TiO2@NCD4h and elemental mapping for Ti, O, C, and N.
Figure 2. (a1a3) FETEM of TiO2@NCD4h. (b) HAADF-STEM of TiO2@NCD4h and elemental mapping for Ti, O, C, and N.
Catalysts 12 01281 g002
Figure 3. XRD patterns of TiO2 and TiO2@NCD2h-8h.
Figure 3. XRD patterns of TiO2 and TiO2@NCD2h-8h.
Catalysts 12 01281 g003
Figure 4. Raman spectra of TiO2 and TiO2@NCD4h.
Figure 4. Raman spectra of TiO2 and TiO2@NCD4h.
Catalysts 12 01281 g004
Figure 5. XPS spectra of pristine TiO2 and TiO2@NCD4h (a) survey scan (b) Ti 2p (c) C 1s (d) N 1s.
Figure 5. XPS spectra of pristine TiO2 and TiO2@NCD4h (a) survey scan (b) Ti 2p (c) C 1s (d) N 1s.
Catalysts 12 01281 g005
Figure 6. Absorption spectrum of (a) NCDs solutions. (b) Pristine TiO2 and TiO2@NCDs thin films (c) bandgap energy of Pristine TiO2 and TiO2@NCDs thin films.
Figure 6. Absorption spectrum of (a) NCDs solutions. (b) Pristine TiO2 and TiO2@NCDs thin films (c) bandgap energy of Pristine TiO2 and TiO2@NCDs thin films.
Catalysts 12 01281 g006
Figure 7. (a) Photocurrent density vs. applied potential curves; (b) Transient photocurrent density curves at 1.89 V vs. RHE of the as-prepared photoelectrodes; (c) Stability test of pristine TiO2 and TiO2@NCD4h at 1.89 V vs. RHE.
Figure 7. (a) Photocurrent density vs. applied potential curves; (b) Transient photocurrent density curves at 1.89 V vs. RHE of the as-prepared photoelectrodes; (c) Stability test of pristine TiO2 and TiO2@NCD4h at 1.89 V vs. RHE.
Catalysts 12 01281 g007
Figure 8. (a). Incident photon-to-current conversion efficiency (IPCE) curves; (b) Applied bias ABPE of TiO2, TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h measured at 1.23V vs. RHE.
Figure 8. (a). Incident photon-to-current conversion efficiency (IPCE) curves; (b) Applied bias ABPE of TiO2, TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h measured at 1.23V vs. RHE.
Catalysts 12 01281 g008
Figure 9. (a) EIS spectra; (b) 0 Mott–Schottky plots of TiO2, TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h. (c) Schematic energy levels of TiO2, TiO2@NCD4h.
Figure 9. (a) EIS spectra; (b) 0 Mott–Schottky plots of TiO2, TiO2@NCD2h, TiO2@NCD4h, TiO2@NCD6h, and TiO2@NCD8h. (c) Schematic energy levels of TiO2, TiO2@NCD4h.
Catalysts 12 01281 g009
Scheme 1. Schematic illustration of TiO2@NCDs.
Scheme 1. Schematic illustration of TiO2@NCDs.
Catalysts 12 01281 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Thanh Thuy, C.T.; Shin, G.; Jieun, L.; Kim, H.D.; Koyyada, G.; Kim, J.H. Self-Doped Carbon Dots Decorated TiO2 Nanorods: A Novel Synthesis Route for Enhanced Photoelectrochemical Water Splitting. Catalysts 2022, 12, 1281. https://doi.org/10.3390/catal12101281

AMA Style

Thanh Thuy CT, Shin G, Jieun L, Kim HD, Koyyada G, Kim JH. Self-Doped Carbon Dots Decorated TiO2 Nanorods: A Novel Synthesis Route for Enhanced Photoelectrochemical Water Splitting. Catalysts. 2022; 12(10):1281. https://doi.org/10.3390/catal12101281

Chicago/Turabian Style

Thanh Thuy, Chau Thi, Gyuho Shin, Lee Jieun, Hyung Do Kim, Ganesh Koyyada, and Jae Hong Kim. 2022. "Self-Doped Carbon Dots Decorated TiO2 Nanorods: A Novel Synthesis Route for Enhanced Photoelectrochemical Water Splitting" Catalysts 12, no. 10: 1281. https://doi.org/10.3390/catal12101281

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop